ArticlePDF Available

Swelling-induced long-range ordered structure formation in polyelectrolyte hydrogel

Authors:

Abstract and Figures

A millimeter-scale periodic structure is created in a polyelectrolyte hydrogel by the rapid-heterogeneous swelling process, and is frozen by the polyion complexation of the polyelectrolyte network with the oppositely charged, semi-rigid polyelectrolyte. The hydrogel is synthesized from a cationic monomer, N-[3-(N,N-dimethylamino)propyl] acrylamide methyl chloride quaternary (DMAPAA-Q), in the presence of a small amount of the oppositely charged poly(2,2′-disulfonyl-4,4′-benzidine terephthalamide) (PBDT) that has a semi-rigid nature. During the swelling process, surface creasing due to the large mismatching of swelling degree between the surface layer and the inner one of the poly DMAPAA-Q (PDMAPAA-Q) gel occurs, which induces highly oriented semi-rigid PBDT molecules along the tensile direction of the crease pattern. To accompany the evolution of surface creasing, a lattice-like periodic birefringence pattern is formed, which is frozen permanently by the strong polyion complex formation, even after the surface instability pattern of the gel disappears completely throughout the dynamic coalescence. In this work we rationally clarified that formation of such a long-range ordered non-equilibrium structure in the polyelectrolyte hydrogel by the rapid-heterogeneous swelling process requires the following three indispensable conditions: (i) swelling-induced surface creasing; (ii) polyion complex formation; and (iii) a semi-rigid or rigid dopant. This sort of non-equilibrium structure formation mechanism may help understand how biomacromolecules that are rigid polyelectrolytes, such as deoxyribonucleic acid, microtubules and actin filaments, form rich architectures during the growth of biological organs.
Content may be subject to copyright.
ISSN 1744-683X
www.rsc.org/softmatter Volume 8 | Number 31 | 21 August 2012 | Pages 7991–8242
Volume 8 | Number 31 | 2012 Soft Matter Themed issue: Hydrogel mechanics Pages 7991–8242
www.rsc.org/softmatter
Registered Charity Number 207890
Highlighting joint research results from the
Department of Chemistry and Biotechnology,
Yokohama National University, Japan, and the
Department of Chemistry and Department of
Chemical Engineering and Materials science,
University of Minnesota, Minnesota.
Title: Thermoreversible high-temperature gelation of an
ionic liquid with poly(benzyl methacrylate-b-methyl
methacrylate-b-benzyl methacrylate) triblock copolymer
A novel thermosensitive triblock copolymer and ionic liquid
composite exhibits a reversible low-temperature-sol–high-
temperature-gel transition.
As featured in:
See Masayoshi Watanabe et al.,
Soft Matter, 2012, 8, 8067.
Themed issue: Hydrogel mechanics
1744-683X(2012)8:31;1-A
PAPER
Jian Ping Gong et al.
Swelling-induced long-range ordered structure formation in
polyelectrolyte hydrogel
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
/ Journal Homepage
/ Table of Contents for this issue
Swelling-induced long-range ordered structure formation in polyelectrolyte
hydrogel
Md. Arifuzzaman,
a
Zi Liang Wu,
a
Takayuki Kurokawa,
bc
Akira Kakugo
bd
and Jian Ping Gong
*
bc
Received 6th April 2012, Accepted 10th May 2012
DOI: 10.1039/c2sm25814e
A millimeter-scale periodic structure is created in a polyelectrolyte hydrogel by the rapid-heterogeneous
swelling process, and is frozen by the polyion complexation of the polyelectrolyte network with the
oppositely charged, semi-rigid polyelectrolyte. The hydrogel is synthesized from a cationic monomer,
N-[3-(N,N-dimethylamino)propyl] acrylamide methyl chloride quaternary (DMAPAA-Q), in the
presence of a small amount of the oppositely charged poly(2,2
0
-disulfonyl-4,4
0
-benzidine
terephthalamide) (PBDT) that has a semi-rigid nature. During the swelling process, surface creasing due
to the large mismatching of swelling degree between the surface layer and the inner one of the poly
DMAPAA-Q (PDMAPAA-Q) gel occurs, which induces highly oriented semi-rigid PBDT molecules
along the tensile direction of the crease pattern. To accompany the evolution of surface creasing,
a lattice-like periodic birefringence pattern is formed, which is frozen permanently by the strong
polyion complex formation, even after the surface instability pattern of the gel disappears completely
throughout the dynamic coalescence. In this work we rationally clarified that formation of such a long-
range ordered non-equilibrium structure in the polyelectrolyte hydrogel by the rapid-heterogeneous
swelling process requires the following three indispensable conditions: (i) swelling-induced surface
creasing; (ii) polyion complex formation; and (iii) a semi-rigid or rigid dopant. This sort of non-
equilibrium structure formation mechanism may help understand how biomacromolecules that are
rigid polyelectrolytes, such as deoxyribonucleic acid, microtubules and actin filaments, form rich
architectures during the growth of biological organs.
Introduction
Biological soft tissues contain a certain amount of water that
ensures molecular mobility and at the same time, possess
a sophisticated structure that enables them to exhibit
outstanding performance over a wide range of physiological
functions.
1–4
Among all, intricate structural patterns found in the
human brain, the intestine, and leaves, are believed to be formed
via a non-equilibrium, dynamic process during growth,
5,6
and are
frozen by the physical/chemical interaction among the mole-
cules.
7,8
Elucidation of the mechanism of such structure
formation induced by non-equilibrium chemistry in living bodies
is an attractive research topic. Furthermore, introducing
sophisticated structures into hydrogels,
9–12
which are soft, wet
materials similar to the biological tissues, by the non-equilibrium
process is one of the ultimate challenges for polymer scientists.
In our previous study, we found that a piece of sheet-like
polyelectrolyte gel containing an oppositely charged semi-rigid
polymer exhibits a periodic birefringence pattern after full
swelling in water.
13
Such a well-ordered structure with milli-
meter-scale periodicity was initially formed in a polyelectrolyte
gel with a dilute concentration (water content >90 wt%).
Although some plausible explanation for the structure formation
could be derived, the exact structure formation mechanism
remains a mystery.
In this study, we showed that the periodic birefringence
pattern, representing an ordered orientation of the semi-rigid
poly(2,2
0
-disulfonyl-4,4
0
-benzidine terephthalamide) (PBDT)
molecules, developed during swelling of the hydrogel. Further,
we found that the surface mechanical instability induced by the
rapid-heterogeneous swelling, the semi-rigid nature of the dopant
macromolecule, and polyion complex formation are the three
indispensable requirements for the formation of such an ordered
structure. This study reveals a novel strategy for introducing
a
Laboratory of Soft and Wet Matter, Division of Biological Sciences,
Graduate School of Science, Hokkaido University, Sapporo 060-0810,
Japan
b
Laboratory of Soft and Wet Matter, Faculty of Advanced Life Science,
Graduate School of Science, Hokkaido University, Sapporo 060-0810,
Japan. E-mail: gong@mail.sci.hokudai.ac.jp
c
Creative Research Institution (CRIS), Hokkaido University, Sapporo
001-0021, Japan
d
PRESTO, Japan Science and Technology Agency (JST), Japan
Electronic supplementary information (ESI) available: Brief discussion
on (i) the POM observations of the as-prepared state, (ii) experimental
results by the controlled (slow) swelling kinetics, and (iii) movie files
recorded during rapid swelling are presented in the ESI. See DOI:
10.1039/c2sm25814e
8060 | Soft Matter, 2012, 8, 8060–8066 This journal is ª The Royal Society of Chemistry 2012
Dynamic Article Links
C
<
Soft Matter
Cite this: Soft Matter, 2012, 8, 8060
www.rsc.org/softmatter
PAPER
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
a long-range ordered structure in amorphous hydrogels through
dynamic mechano-complexation coupling in a non-equilibrium
process. This strategy can be applied to soft, wet systems in
a variety of contexts, namely, for developing various ordered
structures with macroscale correlation by proper designing of the
hydrogel geometry. Furthermore, rigid or semi-rigid macro-
molecules such as deoxyribonucleic acid (DNA), microtubules
(MTs), and actin filaments (F-actin) are ubiquitous in living
organisms. These biomacromolecules form rich architectures
through self-assembly or via physical/chemical interactions with
other species. The mechanism of non-equilibrium structure
formation elucidated in this study might provide scientific
insights into the formation of rich architectures by these bio-
macromolecules during the growth of biological organs.
Experimental
Materials
PBDT, a water-soluble semi-rigid polyanion, was synthesized by
an interfacial polycondensation reaction.
14
The synthesized
PBDT had an average molecular weight, M
w
, of about 1.6
10
6
g mol
1
and a liquid crystalline critical concentration C
LC
*of
2.8 wt% in aqueous solution.
13,15,16
The cationic monomer,
N-[3-(N,N-dimethylamino)propyl]acrylamide methyl chloride
quaternary (DMAPAA-Q) (Kohjin Co. Ltd. Japan), the anionic
monomer, 2-acrylamido-2-methylpropanesulfonic acid (AMPS)
(Tokyo Kasei Co. Ltd.), the neutral monomer, acrylamide
(AAm) (Junsei Chemical Co. Ltd. Japan), and the photoinitiator,
2-oxoglutaric acid (OA) (Wako Pure Chemical Industries Ltd.
Japan), were used as received without further purification. N,N
0
-
Methylenebis(acrylamide) (MBAA) (Wako Pure Chemical
Industries Ltd. Japan) was recrystallized from ethanol and used
as a chemical cross-linker. For all the experiments, deionized
water, purified using 0.22 mm and 5 mm membrane filters, was
used.
Synthesis of the sheet-like polyelectrolyte gels
To synthesize the poly DMAPAA-Q (PDMAPAA-Q) gel con-
taining PBDT, an aqueous solution was prepared by mixing the
cationic monomer, DMAPAA-Q (2.0 M), the semi-rigid anionic
polymer, PBDT (1.0 wt%), the chemical cross-linker, MBAA
(2.0 mol%), and the photoinitiator, OA (0.15 mol%) together (the
amount in mol% is related to the cationic monomer concentra-
tion). After proper mixing, the precursor solution was poured
into a reaction cell consisting of a 1.0 mm thick rectangular
silicone rubber frame sandwiched between a pair of parallel glass
plates. Before polymerization, the precursor solution was trans-
parent and completely amorphous because the PBDT concen-
tration was much lower than the liquid crystalline critical
concentration (C
LC
* ¼ 2.8 wt%). In the precursor solution the
charge ratio of PBDT to DMAPAA-Q was 0.02, which was far
below the stoichiometric value. Radical polymerization was
effected by UV irradiation from both sides of the glass reaction
cell for 6.0 h at room temperature in argon atmosphere. Other
gels with sheet-like shapes [poly AMPS (PAMPS) and poly AAm
(PAAm) gels in the presence of the semi-rigid anionic polymer
PBDT and PDMAPAA-Q gel in the presence of the flexible
anionic polymer PAMPS] were synthesized by using the same
method and conditions.
Swelling of the polyelectrolyte gels
After prolonged UV polymerization, the sheet-like gels (about
45 45 1.0 mm
3
) were carefully removed from the glass
reaction cell and cut into specific dimensions (about 6.0 3.0
1.0 mm
3
) using a mechanical gel cutter (Dumb Bell Co. Ltd.
Japan). Then, the gels were immediately immersed in water for
spontaneous swelling. The volume of water (200 mL) was kept
constant for every sample. In the case of slower kinetics, to
decrease the rate of swelling, a confined metal chamber in which
equilibrated humid conditions were successfully maintained for
a long time was used.
Measurement of swelling kinetics
At room temperature, a set of as-prepared gels with dimensions
of about 6.0 3.0 1.0 mm
3
were immersed together in a large
quantity of water for swelling. The increased mass of the samples
at different swelling times (t) was recorded by measuring the
weight of the samples on an electronic balance (Shimadzu Co.,
Kyoto, Japan). The relative swelling ratio, q (g/g), is defined as
the ratio of the weight of the gel swollen for different time lengths
to its weight in the as-prepared state.
17
Polarizing optical microscope (POM) observation
The surface morphology and the birefringence of the gels in the
as-prepared state, during swelling, and finally in the equilibrium
swelling state were observed under a polarizing optical micro-
scope (POM) (Nikon, Eclipse, LV100POL) in the parallel and
crossed polarization modes, respectively, at room temperature.
To determine the orientation direction of the rod-like PBDT
molecules or their fibrous bundles, the samples were also
observed under a crossed POM with a 530 nm sensitive tint plate.
The change in the surface morphology of the gels during unre-
stricted and rapid swelling was observed under the POM with
a parallel polarizer at room temperature. To observe such
a swelling-induced surface instability crease pattern, the surface
of the as-prepared gels was fully covered with an aqueous solu-
tion of 0.05 wt% Alcian blue (Wako Pure Chemical Industries
Ltd. Japan). Therefore, the crease pattern with its sharp
boundary was clearly observed in the gel during swelling. Images
were captured by a digital camera coupled with the microscope.
Results and discussion
To elucidate the correlation between the surface creasing and the
semi-rigid molecular orientation at different swelling times in
water, we observed the sample under a polarizing optical
microscope (POM) in three different modes: a parallel polari-
zation mode to identify the surface morphology [Fig. 1, column
(i)], a crossed polarization mode to identify the birefringence
[Fig. 1, column (ii)], and a crossed polarization mode with
a 530 nm sensitive tint plate to identify the orientation of the
semi-rigid PBDT molecules [Fig. 1, column (iii)]. The schematic
orientation of the PBDT molecules is shown in Fig. 1,
column (iv).
This journal is ª The Royal Society of Chemistry 2012 Soft Matter, 2012, 8, 8060–8066 | 8061
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
The as-prepared PBDT-containing PDMAPAA-Q gel was
always transparent with a smooth surface [Fig. 1a(i)] and
exhibited irregular birefringence [Fig. 1a(ii and iii)] that was
almost unchanged with the rotation of the sample on the POM
stage (see ESI, Fig. S1†).
When a sample of the as-prepared gel, 6.0 3.0 1.0 mm
3
in
size, was immersed in water, rapid swelling of the gel commenced
immediately and polygonal patterns appeared on the gel surface,
with clear lines that corresponded to the borders of the crease
patterns [Fig. 1b(i)]. Simultaneously with the appearance of the
polygonal patterns, the entire irregular birefringence pattern
rapidly changed into numerous distinct small domains (about
200–300 mm in length) having leaf-like shapes. These leaf-like
patterns always appeared around the borderlines of the crease
patterns [Fig. 1b(ii) and see ESI, Movie S1†]. The birefringence
of the PDMAPAA-Q gel containing PBDT is related to the
orientation of the semi-rigid PBDT molecules, which is typical of
that of a positive liquid crystal.
15,16,18–21
From the characteristic
birefringence colors in the presence of the 530 nm sensitive tint
plate [Fig. 1b(iii)], we concluded that the PBDT molecules are
aligned perpendicular to the borderlines, as illustrated in
Fig. 1b(iv).
As the swelling time increased, these polygonal surface
patterns coalesced into larger ones ca. 1.0 mm in length
[Fig. 1c(i)], exhibiting even more clear, lattice-like borders cor-
responding to the cubic packing of the convex creases. Consistent
with this surface pattern evolution, the leaf-like domains in the
birefringence image coalesced into larger ellipsoidal domains,
Fig. 1 Time evolution of surface morphology and ordered structure formation for PBDT-containing PDMAPAA-Q hydrogel during swelling in water.
The observation was performed under the polarizing optical microscope (POM). The images presented in column (i), observed under the parallel
polarizer, show the surface morphology of the gel; column (ii), observed under the crossed polarizer, shows the birefringence of the gel; column (iii),
observed under the crossed polarizer with a 530 nm sensitive tint plate, shows the molecular orientation direction. Column (iv) is a schematic illustration
of the orientation of semi-rigid PBDT molecules, where the green single bars represent the weak molecular orientation of PBDT, while the blue and
orange bars indicate strong/highly ordered orientation. To visualize the surface morphology, the as-prepared gel was swollen in water containing a dye,
Alcian blue. The dimensions of the as-prepared gels were about 6.0 mm (L) 3.0 mm (W) 1.0 mm (T). A: analyzer, P: polarizer, X
0
and Z
0
: fast and
slow axes of the tint plate, respectively. All the images are shown on the same scale as in a(i).
8062 | Soft Matter, 2012, 8, 8060–8066 This journal is ª The Royal Society of Chemistry 2012
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
with the borderline as the long axis of the ellipsoid and the PBDT
molecules oriented vertically to this axis [Fig. 1c(iii)]. After
prolonged swelling, the surface crease pattern disappeared along
with the lattice-like borders, and the gel surface became smooth
and flat [Fig. 1d(i)], while the birefringence pattern remained
even after reaching the equilibrium swelling state, keeping with
the high periodicity [Fig. 1d(ii and iii)].
To further confirm the lattice-like symmetry of the oriented
structure, the angle dependence of the finally stable birefringence
pattern was investigated. With every 90
rotation, the same
periodic birefringence pattern appeared, while the pattern
became less regular with a 45
rotation in the clockwise or
anticlockwise direction (see ESI, Fig. S2†). These observations
confirmed that the gel has a periodically ordered structure with
a square-shaped lattice unit, whereupon the PBDT molecules
orient vertically with respect to the lattice boundary. One of the
periodical units in the equilibrium state is highlighted by the
dashed boxes in Fig. 1d(ii and iv). Thus, our previously proposed
structure needs modification.
13
Next, we carried out a more quantitative investigation of the
correlation between the swelling kinetics and the ordered struc-
ture formation in the gel. For a diffusion-limited process, the
characteristic swelling time is determined by the smallest
dimension of the sample, which in the present case is the sample
thickness. The time profile of the swelling ratio q in terms of the
weight change of the gel relative to the as-prepared state is shown
in Fig. 2(a). The gel swelled rapidly and reached its equilibrium
swelling state within 20 min [Fig. 2(a), inset]. Evolution of the
surface morphology and the birefringence pattern after the initial
600 s are shown in Fig. 2(b) and (c), respectively. As shown in
Fig. 2(b), bumped irregular polygons appeared on the surface
during the initial period of the swelling process (30 s). With the
advancement of swelling (200 s), the irregular polygons fused
rapidly to form regular patterns with lattice borderlines. In order
to characterize the coalescence of the surface crease patterns, we
define n
JP
as the density of junction point (the number of
crossover junction points of the creases per square millimeter
area), where at least three bumps coexist. The value of n
JP
decreased very rapidly with time and finally reduced to zero after
a swelling time of about 400 s [Fig. 2(b)].
The leaf-like birefringence pattern appeared at around 60 s
and became regular at around 200 s [Fig. 2(c)] like a symmetric
lattice. The leaf length or the length of the lattice unit L
s
increased gradually with the swelling time [see ESI, Movie S2†],
reaching a constant value of 1.1 mm at around 400 s [Fig. 2(c)],
which was approximately the same time taken for the disap-
pearance of the surface crease pattern.
From the above observations, we conclude that a periodic
birefringence pattern is formed during the swelling process and
that it is correlated to surface creasing. The latter is often tran-
sient in nature during the swelling of hydrogels, especially in the
case of polyelectrolyte hydrogels that have a very high osmotic
swelling pressure.
22–39
The rapid swelling of the surface layer
causes a large difference in the degree of swelling between the
surface and the inner layer of the gel; this spatial mismatch results
in compressive strain in the surface layer and leads to transient
creasing instability in the surface layer.
22–39
Theoretical and
experimental studies indicate that the onset of creasing is only
related to the effective compressive strain experienced by the
surface layer, irrespective of the modulus and thickness of the
mismatched layers, while the characteristic spacing between the
creases increases with the surface layer thickness. The critical
value of compressive strain for the creasing instability, 3
c
,is
reported as 0.35.
22,24,29,35–39
At the initial stage of swelling, the
high osmotic swelling pressure induces rapid swelling to satisfy
3 > 3
c
; thus, creasing instability occurs and polygonal patterns
appear. With the advancement of swelling, the thickness of the
surface swollen layer increases, and this leads to coalescence and
eventual disappearance of the crease patterns at equilibrium
Fig. 2 Time profiles of swelling ratio (q) relative to the as-prepared state
(a), density of junction point (n
JP
) of the crease patterns on the surface
(b), and the length of the lattice unit (L
s
) of the birefringence images (c) of
the PBDT-containing PDMAPAA-Q gel during swelling in water. Inset
graph in (a) represents the identical swelling process of the gel for
a prolonged period. Images in (b) and (c) were observed by a POM with
a parallel and crossed polarizer, respectively. The scale bar in the inset
images represents 500 mm, and all the images are shown on the same scale.
This journal is ª The Royal Society of Chemistry 2012 Soft Matter, 2012, 8, 8060–8066 | 8063
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
swelling because of the disappearance of the swelling mismatch.
It is interesting that in the present case, before disappearing, the
polygonal patterns coalesce to form cubic patterns instead of
hexagonal patterns, which is the previously reported mode at the
instability onset.
22,23
We also found that the size of the millimeter-scale periodicity
increased linearly with the sample thickness but did not change
with the formulation of the gel, such as the concentrations of the
cationic monomer and the semi-rigid anionic polymer in the
precursor solution.
13
These results are in agreement with the
feature of the creasing instability.
35–39
To confirm that the surface creasing originated only from the
rapid swelling of the polyelectrolytic gel PDMAPAA-Q and that
the semi-rigid PBDT did not play a role in the same, we studied
the swelling behavior of the PDMAPAA-Q hydrogel without
PBDT [Fig. 3(a)]. No birefringence was observed in the as-pre-
pared state [Fig. 3a(i)], indicating the amorphous nature of the
PDMAPAA-Q hydrogel. A transient surface crease pattern and
birefringence pattern similar to those for PDMAPAA-Q con-
taining PBDT [Fig. 1c(ii)] were observed at the initial stage of
swelling, although the birefringence pattern was relatively weak
[Fig. 3a(ii)]. However, both the patterns disappeared at equilib-
rium swelling [Fig. 3a(iii)].This result indicated that the periodic
birefringence pattern is induced by the creasing instability of the
PDMAPAA-Q gel, while the strong polyion complexation
between the positively charged PDMAPAA-Q network and the
Fig. 3 Effect of surface instability, polyion complexation, and rigidity of the dopant molecule on the regular-stable birefringence pattern formation in
the hydrogel during the swelling process. The positively charged PDMAPAA-Q gel alone [a(i)] or in the presence of the negatively charged flexible
dopant PAMPS [b(i)] exhibits almost no birefringence in the as-prepared state. Upon swelling, a transient birefringence pattern appears in the gels [a(ii)
and b(ii)] because of the surface creasing instability. However, both the gels become amorphous well ahead of equilibrium swelling [a(iii) and b(iii)]. In
presence of the negatively charged semi-rigid PBDT, the as-prepared anionic PAMPS and neutral PAAm gels exhibit weak birefringence patterns [c(i)
and d(i), respectively]. During swelling, the birefringence pattern in the PBDT-containing PAMPS gel changes into numerous leaf-like patterns [c(ii)]
because of the strong surface creasing instability [same as the PBDT-containing PDMAPAA-Q gel in Fig. 1c(ii)]. In contrast, the birefringence pattern
remains almost unchanged in the PBDT-containing PAAm gel [d(ii)] since there is no significant surface instability. However, after equilibrium swelling,
both the PAMPS and PAAm gels containing PBDT show completely amorphous characteristics [c(iii) and d(iii), respectively] owing to the absence of
polyion complexation. In all the cases, the dimensions of the as-prepared gels are the same as those in Fig. 1. A: analyzer, P: polarizer. All the images are
shown on the same scale as in a(iii).
8064 | Soft Matter, 2012, 8, 8060–8066 This journal is ª The Royal Society of Chemistry 2012
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
negatively charged semi-rigid molecule PBDT freezes the bire-
fringence pattern. Certainly, polyion complexation is enhanced
by the outward diffusion of the mobile counterions (Na
+
and
Cl
) from the gel to the solvent environment during the swelling
process,
40–42
which stabilizes the molecular orientation inside the
gel. Thus, the intense polyion complexation between PDMA-
PAA-Q and PBDT freezes the periodic molecular orientation
induced by the mechanical instability.
Furthermore, we have clarified that the semi-rigid nature of
PBDT plays an indispensable role in the stabilization of the
birefringence pattern. As shown in Fig. 3b, the PDMAPAA-Q
hydrogel containing a negatively charged flexible linear polymer,
polyAMPS (PAMPS), exhibited no birefringence in the as-
prepared state [Fig. 3b(i)], while the crease pattern and bire-
fringence pattern appeared transiently during swelling
[Fig. 3b(ii)]. However, both patterns disappeared well ahead of
the equilibrium swelling state [Fig. 3b(iii)]. This result suggests
that the semi-rigid nature of the PBDT plays an important role in
maintaining the highly oriented structure in the gel. As has been
elucidated, mesoscopic fibrous bundles formed from semi-rigid
PBDTs are very sensitive to external stress–strain and show
a stronger birefringence owing to specific molecular orientation
than do the flexible molecules.
43
The strong birefringence
observed in Fig. 1 indicates that the semi-rigid PBDT forms long
and rigid fibrous bundles in the tensile direction of creasing,
almost vertical to the borders of the crease patterns.
We also confirmed the effect of polyion complexation by
investigating the like-charged combination. The gel synthesized
from an anionic monomer, AMPS in the presence of the semi-
rigid PBDT also exhibited certain birefringence in the as-
prepared state [Fig. 3c(i)] and an creasing instability pattern
when exposed to water [Fig. 3c(ii)]. However, after sufficient
(equilibrium) swelling, both the surface instability and birefrin-
gence pattern disappeared [Fig. 3c(iii)]. Such an amorphous
phenomenon can be explained by the absence of polyion
complexation between the semi-rigid anionic PBDT and the
chemically cross-linked negatively charged PAMPS. We also
confirmed that in the absence of creasing instability, for example,
in the case of the PDMAPAA-Q gel swollen with water vapor
(see ESI, Fig. S3†) or the neutral PAAm gel, no periodic bire-
fringence pattern appeared, even though these gels contained
PBDT [Fig. 3d]. Table 1 summarizes the results obtained for
various systems.
Conclusion
In conclusion, a long-range periodic structure is induced during
the swelling of the polyelectrolyte gel, concomitantly with the
occurrence of surface creasing, a transient process related to the
mechanical instability during the free and rapid swelling of
polyelectrolytic gels in water. The mismatch in the swelling ratio
between the surface layer and the inner layer of the gel induces
creasing instability, which exerts tension on the semi-rigid
molecules vertical to the borderline of the crease patterns. Thus,
the molecules orient along the tensile direction and exhibit strong
Table 1 Surface morphology and birefringence pattern of various hydrogels with different dopant molecules. A swelling-induced permanent and highly
ordered macroscopic structure is obtained only for the PDMAPAA-Q gel containing PBDT. Here, the circles and crosses indicate ‘yes’ and ‘no,’
respectively. The symbols (+), (), and (0) indicate positively charged, negatively charged, and neutral polymers, respectively
Hydrogel
Polyelectrolyte
dopant Crease pattern
Lattice-like birefringence
During swelling At equilibrium
PDMAPAA-Q (+) PBDT() BB B
PBDT()
a

PAMPS() BB
None BB
PAMPS() PBDT() BB
PAAm(0) PBDT() 
a
Slow swelling in water vapor.
Fig. 4 Mechanism of the formation of swelling-induced ordered struc-
ture in a polyelectrolyte gel containing a very small amount of an oppo-
sitely charged semi-rigid polyelectrolyte. When the sheet-like gel is
immersed in water, rapid swelling occurs in the surface layer owing to the
high osmotic pressure of the polyelectrolytic gel. The rapid swelling of the
surface layer induces a large mismatch in the internal stress between the
surface layer and the inner layer. Therefore, surface creasing instability
occurs, which exerts tensile stress vertical to the boundary of the crease
patterns and induces polymer network orientation. In the presence of an
oppositely charged semi-rigid polyelectrolyte as the dopant, this polymer
network orientation is frozen by the subsequent polyion complexation
and fibrous bundle formation between the two oppositely charged
components, which is stabilized by the removal of the low-molecular
counterions through diffusion from the gel. Thus, the ordered structure
remains even after the surface crease pattern completely disappears in the
equilibrium swelling state. The dotted red arrows indicate the direction of
tension induced by the formation of the crease pattern.
This journal is ª The Royal Society of Chemistry 2012 Soft Matter, 2012, 8, 8060–8066 | 8065
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
periodic birefringence. The molecular orientation is fixed
simultaneously by the strong ion complexation between the two
oppositely charged components and the formation of a lateral
fibrous bundle from these components; the semi-rigid nature of
the dopant molecules is indispensable for the fibrous bundle
formation. Thus, the strong birefringence pattern is frozen even
after the disappearance of the surface creasing instability. Fig. 4
illustrates the mechanism of formation of the swelling-induced
ordered structure in the polyelectrolyte gel.
These results might help derive a novel strategy for introducing
an ordered structure in soft, wet systems using rigid macromol-
ecules, coupling with mechanical instability and ion
complexation.
Acknowledgements
This research was supported by a Grant-in-Aid for the Specially
Promoted Research (no. 18002002) from the Ministry of
Education, Science, Sports and Culture of Japan. The authors
are thankful to Mr Yukihiro Nakano for informative and valu-
able discussions.
References
1 J. F. V. Vincent, in Structural Biomaterials, rev. edn, Princeton
University Press, Princeton, New Jersey, USA, 1991.
2 M. Spencer, W. Fuller, M. H. F. Wilkins and G. L. Brown, Nature,
1962, 194, 1014.
3 C. Sanchez, H. Arribart and M. M. Giraud-Guille, Nat. Mater., 2005,
4, 277.
4 Hierarchical Structures in Biology as a Guide for New Materials
Technology, ed. D. A. Tirrell, National Material Advisory Board,
The National Academic Press, Washington DC, 1994.
5 E. Hohlfeld and L. Mahadevan, Phys. Rev. Lett., 2011, 106, 105702.
6 J. Yin, G. Gerling and X. Chen, Acta Biomater., 2010, 6, 1487.
7 S. I. Stupp, V. LeBonheur, K. Walker, L. S. Li, K. E. Huggins,
M. Keser and A. Amstutz, Science, 1997, 276, 384.
8 J. M. Lehn, Supramolecular Chemistry: Concepts and Perspectives,
Wiley-VCH, Weinheim, Germany, 1995.
9 Y. Osada and A. Matsuda, Nature, 1995, 376, 219.
10 M. Hayakawa, T. Onda, T. Tanaka and K. Tsujii, Langmuir, 1997,
13, 3595.
11 T. Kato, N. Mizoshita and K. Kishimoto, Angew. Chem., Int. Ed.,
2006, 45, 38.
12 T. Dobashi, K. Furusawa, E. Kita, Y. Minamisawa and
T. Yamamoto, Langmuir, 2007, 23, 1303.
13 Z. L. Wu, Md. Arifuzzaman, T. Kurokawa, H. Furukawa and
J. P. Gong, Soft Matter, 2011, 7, 1884.
14 E. J. Vandenberg, W. R. Diveley, L. J. Filar, S. R. Pater and
H. G. Barth, J. Polym. Sci., Part A: Polym. Chem., 1989, 27, 3745.
15 T. Funaki, T. Kaneko, K. Yamaoka, Y. Oshedo, J. P. Gong,
Y. Osada, Y. Shibasaki and M. Ueda, Langmuir, 2004, 20, 6518.
16 Y. Shigekura, Y. M. Chen, H. Furukawa, T. Kaneko, D. Kaneko,
Y. Osada and J. P. Gong, Adv. Mater., 2005, 17, 2695.
17 H. Furukawa, J. Mol. Struct., 2000, 554, 11.
18 Z. L. Wu, T. Kurokawa, S. M. Liang, H. Furukawa and J. P. Gong, J.
Am. Chem. Soc., 2010, 132, 10064.
19 D. Demus, J. Goodby, G. W. Gray, H. W. Spiess, and V. Vill,
Handbook of Liquid Crystals, Wiley-VCH, Weinheim, Germany,
1998.
20 Z. L. Wu, D. Sawada, T. Kurokawa, A. Kakugo, W. Yang,
H. Furukawa and J. P. Gong, Macromolecules, 2011, 44, 3542.
21 K. Murata and K. Haraguchi, J. Mater. Chem., 2007, 17, 3385.
22 T. Tanaka, S.-T. Sun, Y. Hirokawa, S. Katayama, J. Kucera,
Y. Hirose and T. Amiya, Nature, 1987, 325, 796.
23 T. Tanaka, Phys. A, 1986, 140, 261.
24 J. W. Kim, J. W. Woon and R. C. Hayward, Nat. Mater., 2010, 9, 159.
25 T. Mora and A. Boudaoud, Eur.Phys. J. E, 2006, 20, 119.
26 H. Tanaka and T. Sigehuzi, Phys. Rev. E: Stat. Phys., Plasmas, Fluids,
Relat. Interdiscip. Top., 1994, 49, R39.
27 H. Tanaka, H. Tomota, A. Takasu, T. Hayashi and T. Nishi, Phys.
Rev. Lett. , 1992, 68, 2794.
28 M. Guvendiren, S. Yang and J. A. Burdick, Adv. Funct. Mater., 2009,
19, 3038.
29 A. N. Gent and I. S. Cho, Rubber Chem. Technol., 1999, 72, 253.
30 E. Sultan and A. Boudaoud, J. Appl. Mech., 2008, 75, 051002-1.
31 S. Yang, K. Khare and P. C. Lin, Adv. Funct. Mater., 2010, 20, 2550.
32 H. S. Kim and A. J. Crosby, Adv. Mater., 2011, 23, 4188.
33 K. Saha, J. Kim, E. Irwin, J. Yoon, F. Momin, V. Trujillo,
D. V. Schaffer, K. E. Healy and R. C. Hayward, Biophys. J., 2010,
99, 94.
34 M. K. Kang and R. Huang, J. Mech. Phys. Solids, 2010, 58, 1582.
35 V. Trujillo, J. Kim and R. C. Hayward, Soft Matter, 2008, 4, 564.
36 M. Guvendiren, J. A. Burdick and S. Yang, Soft Matter, 2010, 6,
5795.
37 L. Jin, S. Q. Cai and Z. G. Suo, Europhys. Lett., 2011, 95, 64002.
38 S. Q. Cai, D. Chen, Z. G. Suo and R. C. Hayward, Soft Matter, 2012,
8, 1301.
39 W. Hong, X. H. Zhao and Z. G. Suo, Appl. Phys. Lett., 2009, 95,
111901.
40 K. Wagner, D. Harries, S. May, V. Kahl, J. O. R
adler and A. Ben-
Shaul, Langmuir, 2000, 16, 303.
41 F. Bordi, C. Cametti, M. Diociaiuti, D. Guadino, T. Gili and
S. Sennato, Langmuir, 2004, 20, 5214.
42 J. Gummel, F. Cousin and F. Bou
e, J. Am. Chem. Soc., 2007, 129,
5806.
43 Z. L. Wu, T. Kurokawa, D. Sawada, J. Hu, H. Furukawa and
J. P. Gong, Macromolecules, 2011, 44, 3535.
8066 | Soft Matter, 2012, 8, 8060–8066 This journal is ª The Royal Society of Chemistry 2012
Published on 06 June 2012. Downloaded by Zhejiang University on 16/10/2013 06:11:12.
View Article Online
... [6] Since water is a critical component of hydrogels, one would reasonably regulate the functions of hydrogels by tuning water contents [7][8][9] and controlling water transportation. [10][11][12][13] However, it is still extremely challenging to manipulate the movement of water in artificial hydrogels. [14] The water transportation is largely affected by polymer hydrophilicity and hydrogel microstructures, which are strongly affected by the hydrogel composition and the preparation method. ...
Article
Full-text available
Controlling water transportation within hydrogels makes hydrogels attractive for diverse applications, but it is still a very challenging task. Herein, a novel type of dually electrostatically crosslinked nanocomposite hydrogel showing thermoresponsive water absorption, distribution, and dehydration processes are developed. The nanocomposite hydrogels are stabilized via electrostatic interactions between negatively charged poly(acrylic acid) and positively charged layered double hydroxide (LDH) nanosheets as well as poly(3-acrylamidopropyltrimethylammonium chloride). Both LDH nanosheets as crosslinkers and the surrounding temperatures played pivotal roles in tuning the water transportation within these nanocomposite hydrogels. By changing the surrounding temperature from 60 to 4 °C, these hydrogels showed widely adjustable swelling times between 2 and 45 days, while the dehydration process lasted between 7 and 27 days. A swift temperature decrease, for example, from 60 to 25 °C, generated supersaturation within these nanocomposite hydrogels, which further retarded the water transportation and distribution in hydrogel networks. Benefiting from modified water transportation and rapidly alternating water uptake capability during temperature change, pre-loaded compounds can be used to track and visualize these processes within nanocomposite hydrogels. At the same time, the discharge of water and loaded compounds from the interior of hydrogels demonstrates a thermoresponsive sustained release process.
... For example, hydrogels based on organic crosslinked polyacrylamide or nanoclaycrosslinked poly(N-isopropylacrylamide) have been successfully used for fabricating underwater superoleophobic surfaces [20][21][22]. However, the swelling nature and salt instability of hydrogels in seawater may damage the original micro/nanostructures on their surfaces [23][24][25], undermining the underwater superoleophobicity [26]. Otherwise, synthetic underwater superoleophobic hydrogels generally suffer from poor mechanical properties in marine environment [27]. ...
Article
The accumulation of microorganisms, algae, mussels, and barnacles on the hull of sea vessels leads to biofouling, surface corrosion, and increased drag as the vessel moves through water costing the marine industry around $15 billion/year. Current commercial traditional antifouling coatings suffer from reduced energy efficiency and short lifespan and contain heavy metals that are toxic to marine organisms and humans. The banning of these coatings is due to environmental concerns, and nonbiocidal alternatives such as polymer-based coatings are being sought after. This review demonstrates emerging promises of hydrogel fouling-release coatings (FRCs) in the marine environment that may be effective against a prevalent amount of biofouling agents. The review also highlights the importance of polymer backbone materials with surface wettability characteristics possessing hydrophobic, hydrophilic, and zwitterionic properties and further discusses emerging antifouling techniques, synthesis of hydrogel, swelling behavior of hydrogel, laboratory assays of hydrogel coating, marine field test, and outlook of hydrogel fouling-release (FRCs) coatings.
Article
Shape memory polymers (SMPs) with multiple functionalities have great potential in implantable biomedical devices, especially vascular stents. However, stents made of SMPs are generally faced with the problem of insufficient radial support due to the sharp decline of the modulus after shape recovery. Therefore, it is necessary to improve the modulus of SMPs after opening the narrow part by other means. In this study, the novel SMPs available for vascular stents were developed with impressive water-induced stiffening when shape recovered in a physiological environment. Herein, a series of shape memory polyurethanes (SMPUs) containing full hard segments on the main chains and bearing hydrophilic tertiary amine soft segments on the side chains were synthesized. When immersed in water, the soft segments were dramatically separated from the hard segments, which were aggregated more to form densely packed hard domains with stronger hydrogen bonding and higher crystallinity. Both Young's modulus and the shape recovery ratio were thus promoted due to the segmental rearrangement in water. At the same time, hydrophilic side chains migrated to the surface driven by the segmental rearrangement in water, which promotes the adhesion and growth of vascular endothelial cells and inhibits the activation of the coagulation system. The ingenious structural design provided SMPUs with adequate mechanical strength and hemocompatibility to qualify for potential applications in self-expanding vascular stents.
Article
Energy and water are of fundamental importance for our modern society, and advanced technologies on sustainable energy storage and conversion as well as water resource management are in the focus of intensive research worldwide. Beyond their traditional biological applications, hydrogels are emerging as an appealing materials platform for energy- and water-related applications owing to their attractive and tailorable physiochemical properties. In this review, we highlight the highly tunable synthesis of various hydrogels, involving key synthetic elements such as monomer/polymer building blocks, cross-linkers, and functional additives, and discuss how hydrogels can be employed as precursors and templates for architecting three-dimensional frameworks of electrochemically active materials. We then present an in-depth discussion of the structure-property relationships of hydrogel materials based on fundamental gelation chemistry, ultimately targeting properties such as enhanced ionic/electronic conductivities, mechanical strength, flexibility, stimuli-responsiveness, and desirable swelling behavior. The unique interconnected porous structures of hydrogels enable fast charge/mass transport while offering large surface areas, and the polymer-water interactions can be regulated to achieve desirable water retention, absorption, and evaporation within hydrogels. Such structure-derived properties are also intimately coordinated to realize multifunctionality and stability for different target devices. The plethora of stimulating examples is expounded with a focus on batteries, supercapacitors, electrocatalysts, solar water purification, and atmospheric water harvesting, which showcase the unprecedented technological potential enabled by hydrogels and hydrogel-derived materials. Finally, we study the challenges and potential ways of tackling them to reveal the underlying mechanisms and transform the current development of hydrogel materials into sustainable energy and water technologies.
Article
This study demonstrates that the bulk alignment of chromonic aggregates can be achieved during the swelling of hydrogels. Swelling of an ionic hydrogel immersed in an aqueous solution of disodium cromoglycate reorients the chromonic aggregates, and millimeter‐thick optically anisotropic hydrogels are obtained. These anisotropic hydrogels contain the chromonic aggregates at a condensed concentration as high as in the columnar phase of a normal chromonic aqueous solution, although the X‐ray diffraction results show much less stacking order and orientational order of the aggregates. Furthermore, anisotropic mechanical properties of the hydrogels are observed due to the anisotropic alignment of the chromonic aggregates. Disodium cromoglycate chromonic aggregates are incorporated in a positively charged hydrogel through ionic bonds. They are uniaxially aligned during the swelling of the hydrogel. Such chromonic‐aggregates‐incorporated hydrogels are optically anisotropic and show anisotropic mechanical properties during stretching.
Article
We prepared poly(N-isopropylacrylamide-r-N-3-(aminopropyl)methacrylamide) (poly(NIPAAm-r-NAPMAm)) gels with poly NIPAAm (PNIPAAm) grafted only in the surface region (so-called thermoresponsive “surface-grafted gels”) with various graft densities, and investigated the effect of the graft density on the bulk properties, such as shrinking and swelling, in response to temperature changes. Initiators for atom transfer radical polymerization (ATRP) and structurally analogous compounds were introduced at certain ratios to the surface regions of the gels, and a subsequent activator regenerated by electron transfer ATRP of NIPAAm was conducted in aqueous media. The graft densities and molecular weights of the grafted polymers were evaluated from the increment in the dry mass of the gels and the amount of introduced ATRP initiators which was measured by elemental analyses. 3D measuring laser microscopy revealed that the prepared gels had graft density-dependent fine wrinkle structures on their surfaces. The surface-grafted gels induced the formation of skin layers during the shrinking process in response to a temperature increase, and their permeability strongly depended on the graft density. The graft density also controlled the kinetics of the swelling behavior in response to a temperature decrease. These physical properties were discussed based on the Young’s modulus of the surface measured by an atomic force microscopy force curve measurement and the homogeneity of the surface polymer network observed by cryo-scanning electron microscopy. This makes it possible to arbitrarily control the characteristics of gels as “open” or “semi-closed” systems, which was uniquely determined by the designs of the surface gel networks.
Article
A macroscopic silk-like fiber can be drawn directly from a hydrogel self-assembled from peptides, which offers an easy and versatile method for generating materials with hierarchically ordered structures over different scales and extends the utility of peptide hydrogels.
Article
Full-text available
We explored the energy transfer dynamics of chlorophyll-aentrapped polyacrylamide hydrogel with a vision of applying this hybrid material in a bio-inspired light harvesting system. Prominent photocurrent response was observed from a simple photovoltaic assembly prepared by encapsulating the hydrogel within two electrodes. For a better understanding of the energy transport, the hybrid systems were synthesized via two different methods (in situ and swelling induced). The difference in photocurrent efficiency among the two systems could be correlated with the different dynamic behaviors of the various excitonic packets and could be explained with environment-assisted transport of photosynthesis (ENAQT). This theory predicts that energy migration in a natural photosystem relies on the balance between coherence and dephasing, owing to the respective packing geometry. Fluorescence anisotropy supports the swollen induced arrangement of chlorophyll-a packets, which are inter-connected in terms of overlapping in site energy, as evident from the broadening of the UV-Vis spectra wherein coherent spreading occurs among around 4 chlorophyll-a molecules, as revealed from the time correlated single photon count. Exciton localization within the packets is destroyed by low frequency noise to mitigate coherence trapping in a coherent classical intermediate fashion, similar to ENAQT in natural photosynthesis, resulting in greater photocurrent efficiency. Optical and photo-physical properties of the in situ sample show that charged dimers are randomly spread throughout the system. As a result, delocalization is destroyed and energy propagation becomes less effective as the exciton is more likely to recombine than trap. The experimental signature of the environment-assisted energy transfer is further supported by a simple mathematical formulation inspired from Kassal's work. The concept of using swollen induced bio-inspired soft materials for solar energy harvesting can pave the way towards a new class of biomimetic solar cell.
Article
Full-text available
The creasing instability of elastomer films under compression is studied by a combination of experiment and numerical simulation. Experimentally, we attach a stress-free film on a much thicker and stiffer pre-stretched substrate. When the substrate is partially released, the film is uniaxially compressed, leading to formation of an array of creases beyond a critical strain. The profile of the folded surface is extracted using confocal fluorescence microscopy, yielding the depths, spacings, and shapes of creases. Numerically, the onset and development of creases are simulated by introducing appropriately sized defects into a finite-element mesh and allowing the surface of the film to self-contact. The measurements and simulations are found to be in excellent agreement.
Article
Full-text available
We report a novel giant oriented structure observed in plate hydrogels synthesized by photo-polymerization of cationic monomers with a cross-linker in the presence of a semi-rigid polyanion as the dopant. The giant structure, formed via self-assembly of the semi-rigid polyion complex, consists of millimetre-scale cubic packed concentric cylindrical domains that are sandwiched by two homeotropically aligned outer layers. A universal relationship between the diameter of the cylinders D and the thickness of the swollen gel T is observed, as D = 0.5T, regardless the change in the concentrations of the polyanion and precursor cationic monomer. This result permits us to induce the giant concentric structure into hydrogels with tunable cylindrical sizes.
Article
Full-text available
When a block of an elastomer is bent, the compressed surface may form a crease. The critical strain for creasing measured experimentally is known to disagree with that predicted by linear perturbation analysis. This paper calculates the critical strain by comparing the elastic energy in a creased body and that in a smooth body. This difference in energy is expressed by a scaling relation. Critical conditions for creasing are determined for elastomers subject to general loads and gels swelling under constraint. The theoretical results are compared with existing experimental observations.
Chapter
IntroductionTheoretical Prediction of the Biaxial Nematic PhaseStructural FeaturesSynthesisCharacterization Methods Concluding RemarksReferences
Article
Biot1 carried out theoretical analyses of the critical strains at which surface instabilities would be encountered in large blocks of a neo-Hookean material subjected to different degrees of compression in two directions. Three special cases are reviewed here: simple compression in one direction, compression in one direction with the perpendicular direction constrained (pure shear), and equibiaxial compression. In all cases a surface instability is predicted to occur at a moderate compressive strain, ranging from about 33% to about 55%. Because unidirectional compression occurs on the inner surface of a rubber block subjected to simple bending, a similar surface instability would be expected at a critical degree of bending, when the surface compression is about 46%. Observations on bent rubber blocks are compared with this theoretical prediction. Sharp creases occurred on the inner surface at a critical degree of bending but the critical compressive strain there was only about 35% and the bending curvature was less than predicted, only about one-half as severe. The cause of this discrepancy is not known.
Article
Optical anisotropy was observed in nanocomposite hydrogels with polymer–clay network structures and was found to exhibit unique changes when the gels were deformed uniaxially. Birefringence measurements showed distinct maxima and sign inversions that depend on strain and gel composition.
Article
Surface-attached hydrogels provide a convenient means to tune interfacial material properties such as biocompatibility and tribology. When the gel undergoes hydration, however, the substrate provides a constraint against lateral expansion, thereby generating an in-plane compressive stress within the gel. For sufficiently large degrees of compression a creasing instability takes place, in which the gel surface locally buckles and forms sharp folds. While this instability has been known in practice for well over a century, it remains poorly understood. Using model polyacrylamide hydrogel systems, we have studied the onset of creasing as a function of material properties and gel thickness, and addressed basic questions regarding crease morphologies and growth mechanisms. Using the understanding gained from these studies, we are developing this instability as a route to create active surfaces, where both surface topography and chemical patterns can be controllably modulated.
Article
In response to external stimuli, polymeric hydrogels can change volume and shape dramatically. Experimental studies have observed a variety of instability patterns of hydrogels, due to swelling or shrinking, many of which have not been well understood. The present paper considers swell-induced surface instability of a hydrogel layer on a rigid substrate. Based on a recently developed theoretical framework for neutral polymeric gels, a linear perturbation analysis is performed to predict the critical condition for the onset of the surface instability. Using a nonlinear finite element method, numerical simulations are presented to show the swelling process, with the evolution of initial surface perturbations followed by the formation of crease-like surface patterns. In contrast to previously suggested critical conditions for surface creasing, the present study suggests a material-specific condition that predicts a range of critical swelling ratios from about 2.5 to 3.4 and quantitatively relates the critical condition to material properties of the hydrogel system. A stability diagram is constructed with two distinct regions for stable and unstable hydrogels depending on two dimensionless material parameters.
Book
Part 1 From molecular to supramolecular chemistry: concepts and language of supramolecular chemistry. Part 2 Molecular recognition: recognition, information, complementarity molecular receptors - design principles spherical recognition - cryptates of metal cations tetrahedral recognition by macrotricyclic cryptands recognition of ammonium ions and related substrates binding and recognition of neutral moelcules. Part 3 Anion co-ordination chemistry and the recognition of anionic substrates. Part 4 Coreceptor molecules and multiple recognition: dinuclear and polynuclear metal ion cryptates linear recognition of molecular length by ditopic coreceptors heterotopic coreceptors - cyclophane receptors, amphiphilic receptors, large molecular cage multiple recognition in metalloreceptors supramolecular dynamics. Part 5 Supramolecular reactivity and catalysis: catalysis by reactive macrocyclic cation receptor molecules catalysis by reactive anion receptor molecules catalysis with cyclophane type receptors supramolecular metallo-catalysis cocatalysis - catalysis of synthetic reactions biomolecular and abiotic catalysis. Part 6 Transport processes and carrier design: carrier-mediated transport cation-transport processes - cation carriers anion transport processes - anion carriers coupled transport processes electron-coupled transpoort in a redox gradient proton-coupled transport in a pH gradient light-coupled transport processes transfer via transmembrane channels. Part 7 From supermolecules to polymolecular assemblies: heterogeneous molecular recognition - supramolecular solid materials from endoreceptors to exoreceptors - molecular recognition at surfaces molecular and supramolecular morphogenesis supramolecular heterogeneous catalysis. Part 8 Molecular and supramolecular devices: molecular recognition, information and signals - semiochemistry supramolecular photochemistry - molecular and supramolecular photonic devices light conversion and energy transfer devices photosensitive molecular receptors photoinduced electron transfer in photoactive devices photoinduced reactions in supramolecular species non-linear optical properties of supramolecular species supramolecular effects in photochemical hole burning molecular and supramolecular electronic devices supramolecular electrochemistry electron conducting devices - molecular wires polarized molecular wires - rectifying devices modified and switchable molecular wires molecular magnetic devices molecular and supramolecular ionic devices tubular mesophases. (Part contents).
Article
Polymer gels, consisting of a cross-linked polymer network immersed in liquid, undergo a volume phase transition: when external conditions such as temperature or solvent composition change, a gel reversibly swells or shrinks, but does so discontinuously1–5. The volume change at the transition can be as large as a factor of one thousand2, and the phenomenon occurs in all gels6,7. The equilibrium aspects of the phase transition have been extensively studied, but its kinetics have not yet been fully explored. In particular, the appearance of patterns on the originally smooth surface of a gel during the transition makes the kinetic process difficult to understand. Here we elucidate the physical basis underlying the formation and evolution of the pattern.